DOI:
10.1039/C6RA11085A
(Paper)
RSC Adv., 2016,
6, 58925-58932
Self-assembled mesoporous Ni0.85Se spheres as high performance counter cells of dye-sensitized solar cells†
Received
29th April 2016
, Accepted 13th June 2016
First published on 15th June 2016
Abstract
Transition metal chalcogenides with mesoporous structure are of special interest for application in electrocatalytic reactions due to their ample unique properties and functionalities. Nevertheless, relevant research on mesoporous Ni0.85Se applied as electrocatalysts is rare. In this contribution, mesoporous Ni0.85Se spheres consisting of numerous primary particles are synthesized via a facile self-assembly route. Mesoporous Ni0.85Se spheres exhibit higher electrocatalytic activity (PCE = 7.24%) for the reduction of triiodide to iodide and lower charge-transfer resistance (Rct = 2.40 Ω) than Ni0.85Se nanoparticles (PCE = 6.73%, Rct = 3.46 Ω) at the electrolyte–electrode interface in dye-sensitized solar cells (DSSCs). In order to further enhance the overall electrocatalytic performance of mesoporous Ni0.85Se spheres, reduced graphene oxide (RGO) and single wall carbon nanotubes (SWCNTs) are introduced by facile physical mixing. RGO and SWCNT not only improved the charge transfer ability, but also yielded favorable synergistic catalytic effects with mesoporous Ni0.85Se spheres. Simultaneously, the PCE values of Ni0.85Se + 6 wt% RGO and Ni0.85Se + 9 wt% SWCNT reach 7.87% and 7.74%, respectively, which are both higher than Pt (7.56%).
Introduction
Functional nanomaterials with mesoporous nanostructures have recently attracted considerable attention due to their distinctive properties including tunable mesopore structures, large internal surface area and their easily functionalized frameworks with well-defined nanostructures.1–5 These substantial properties enable mesoporous materials with versatile functions, such as adsorption, catalysis, drug delivery as well as energy conversion and storage.3,6–8 In particular, mesoporous materials have exhibited excellent electrocatalytic activity because their high specific surface areas promote the interface contact between electrode and electrolyte, and the mesopores allow liquid electrolyte to easily diffuse into the electrode materials.9–11 In recent years, considerable work has been devoted to fabricate mesoporous nanostructures of transition metal chalcogenides and make use of them as electrocatalysts due to their excellent electrocatalytic properties resulting from their combination of mesoporous structure characteristic and semiconductor framework nature.12–15 Zhang and his co-workers prepared mesoporous Co3S4 nanosheets via the anion exchange reaction, in addition mesoporous Co3S4 nanosheets were found to be highly active and stable for electrocatalytic oxygen evolution reaction.16 Recently, Chen and his co-workers found that 3D Cu-doped CoS porous nanosheet films provided a great large number of active catalytic sites and easy accessibility toward Sn2−/S2− electrolyte solution, leading to high-performance counter cell for quantum-dot-sensitized solar cells.17
Nickel selenides, a member of transition metal chalcogenides with Pauli paramagnetism and metallic conductivity, have exhibited abundant electrical and magnetic properties.18–24 As an important class of nickel selenides, due to its specific stoichiometries, Ni0.85Se owned abundant unsaturated atoms and unique electronic configuration.25 Furthermore, the small difference in electronegativity between Ni (c = 1.92) and Se (c = 2.55) results in the narrow band gap of Ni0.85Se (about 2.0 eV).26,27 All of these outstanding features of Ni0.85Se endow itself with many catalytic active sites, fast electron-transfer channels, as well as beneficial charge and mass transfer in electrocatalytic reaction. Thus Ni0.85Se has shown excellent electrocatalytic performance as counter cells (CEs) of DSSCs.28,29 So far, although several work has been done to study the nanostructures and electrocatalytic property of Ni0.85Se, relevant research about mesoporous Ni0.85Se applied as electrocatalytic materials is rare.30
In this report, we synthesized specific stoichiometric Ni0.85Se with the morphology of mesoporous spheres consisted of numerously primary particles, and mesoporous Ni0.85Se spheres exhibited average pore diameter of 14.03 nm. The electrocatalytic performance of mesoporous Ni0.85Se spheres was measured as CEs of DSSCs. Mesoporous Ni0.85Se spheres exhibited comprehensively better electrocatalytic performance than Ni0.85Se nanoparticles. In order to further enhance the electrocatalytic performance of mesoporous Ni0.85Se spheres CE, different amount of RGO and SWCNT were introduced by facile physical mix. Electrochemical measurements demonstrated that both of RGO and SWCNT not only improved charge transfer ability of CEs, but also yielded favorable synergistic catalytic effect with mesoporous Ni0.85Se spheres.31 More importantly, the PCE values of Ni0.85Se + 6 wt% RGO and Ni0.85Se + 9 wt% SWCNT reached 7.87% and 7.74%, furthermore, Rct values were reduced to 0.68 Ω and 0.74 Ω, respectively.
Experimental section
Synthesis of mesoporous Ni0.85Se spheres
All reagents are of analytic grade and used without further purification. The synthesis of Ni0.85Se was carried out by a simple thermal reaction, and whole experimental flowchart is schematically illustrated in Fig. 1a. In detail, the original molar ratio between Ni(CH3COO)2·4H2O and selenium powder was selected to be 1
:
1. Ni(CH3COO)2 (0.8 g) was dispersed in tetraethylenepentamine (TEPA) (15 ml) and stirred for 5 min, then ethylene glycol (15 ml) was added to the above solution. After stirring for another 5 min, selenium powder (0.254 g) was subsequently placed into the mixture. Lastly, 5 ml hydrazine hydrate was added dropwise. The resulting solution was put into Teflon-lined autoclave of 50 ml capacity and heated at 150 °C for 15 h. Then, the autoclave was allowed to cool to room temperature naturally. The product was washed with water and absolute ethanol to remove impurities, and then dried at 60 °C.
 |
| Fig. 1 Illustration of the synthesis process for mesoporous Ni0.85Se spheres (a). XRD pattern of as-synthesized mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles (b). | |
Characterization of prepared samples
For characterizing the obtained samples, the crystallinity and composition of the samples were characterized by X-ray diffraction (XRD, D/max-2500, JAPAN SCIENCE) with Cu Kα radiation (λ = 1.54056 Å). The morphology of samples was studied by field-emission scanning electron microscopy (FE-SEM, Nanosem 430, FEI). More detailed insight into the microstructure of the sample was given by high-resolution transmission electron microscopy (TEM, Tecnai G2 F20, operating at 200 kV, FEI). The Brunauer–Emmett–Teller (BET) specific surface area was analyzed by the BET equation using a Tristar 3000 nitrogen adsorption apparatus. Fourier transform infrared spectroscopy (FT-IR) was carried on American Nicolet Instrument Co.
Test and characterization of CEs and DSSCs
The fabrication of CEs and DSSCs was listed in ESI.† All the electrochemical measurements were finished with the Zahner IM6 exelectrochemical workstation. Photocurrent–voltage curves were conducted in simulated AM 1.5 illumination (100 mW cm−2, Trusttech CHF-XM-500 W) with a Keithley digital source meter (Keithley 2410, USA). Cyclic voltammetry (CV) was recorded with a three electrode system on the exelectrochemical workstation. Pt was used as the counter electrode, and Ag/AgCl was used as the reference electrode. An solution of 10.0 mM LiI, 1.0 mM I2, and 0.1 M LiClO4 in acetonitrile served as the electrolyte. CV curves were recorded in the range of −0.4 to 1.2 V at a scan rate of 25 mV s−1. Electrochemical impedance spectra (EIS) analysis was conducted at zero bias potential and the impedance data covered a frequency range of 0.1 Hz to 1 MHz. The amplitude of the sinusoidal AC voltage signal was 5 mV. The analyses of the resulting impedance spectra were conducted using the software Zview 2.0. Tafel polarization measurements were employed in a symmetrical dummy cell which was used in the EIS experiments. The electrolyte was as the same of the electrolyte of DSSC. The scan rate was 20 mV s−1, and the voltage range is −1.0 to 1.0 V.
Results and discussion
XRD, SEM, TEM and BJH
As shown in Fig. 1b, the phase purity of typically as-synthesized mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles were characterized by X-ray diffraction (XRD). All the primary diffraction peaks of both patterns agree with the standard data of Ni0.85Se (JCPDS no. 18-0888, vertical line in Fig. 1b). Diffraction signals from impurities were not found in the both patterns, which indicate that only pure Ni0.85Se phases were present in the two samples. The three strongest peaks are at about 33.04°, 44.6°, 50.24°, corresponding to the (101), (102) and (110) crystal faces respectively. XRD patterns of RGO and Fourier transform infrared (FTIR) transmission spectra of SWCNT are shown in ESI, Fig. S1.†
The morphology of the sample was characterized by scanning electron microscopy (SEM) and transmission electron microscopy (TEM). (Fig. 2a–f) SEM image at lower magnification (Fig. 2a) proves the homogeneity of mesoporous Ni0.85Se spheres. As shown in Fig. 2b and c, each sphere is composed of numerous primary particles with size about 50 nm. The exterior surfaces of the microspheres are not clearly smooth but contain an extensive growth of disordered interparticle porosity structure. A representative high-resolution TEM image is shown in Fig. 2f spacing of the fringes is 0.271 nm, which corresponds well to the (101) plane of Ni0.85Se. Energy dispersion spectrum (EDS) profile indicates the existence of three elements: Cu, Ni and Se (inset in the Fig. 2f). SEM image of Ni0.85Se nanoparticles is shown in ESI, Fig. S2.† And more detailed TEM figures of RGO and SWCNT are presented in Fig. S3 and S4.†
 |
| Fig. 2 Typical SEM (a–c), TEM (d, and e), HRTEM (f) and EDS (inset in f) images of mesoporous Ni0.85Se spheres. | |
The nitrogen adsorption–desorption isotherms and porosity of mesoporous Ni0.85Se spheres are presented in Fig. 3. The recorded adsorption and desorption isotherms of mesoporous Ni0.85Se spheres sample display a type III isotherm with a type H3 hysteresis loop (at P/P0 > 0.8), implying the presence of mesopores (2–50 nm in size). The inset in the Fig. 3 shows the corresponding pore size distribution calculated by the Barrett–Joyner–Halenda (BJH) method from the desorption branch, indicating a broad pore size distribution, from several nanometres to about 100 nm. The average pore diameter of mesoporous Ni0.85Se spheres is 14.03 nm and the pore volume is 0.028 cm3 g−1. The foregoing results further confirmed the porosity structure of mesoporous Ni0.85Se spheres, which is beneficial to absorb more active species and reactants on its surface, as well as contain much reaction active sites.32
 |
| Fig. 3 N2 adsorption/desorption isotherm and Barrett–Joyner–Halenda (BJH) pore size distribution plot (inset) of mesoporous Ni0.85Se spheres. | |
Electrocatalytic performance of mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles
Electrochemical impedance spectroscopy (EIS) was carried out using a symmetric cell (CE/electrolyte/CE) consisting of two identical CEs to investigate the property of charge transfer at the CE/electrolyte interface and electrolyte diffusion in bulk solution or the film pores. As shown in Fig. 4, the EIS of mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles CEs have been investigated, in which an equivalent circuit diagram is also provided for fitting the Nyquist plots by the Z-view software, and the relevant EIS parameters are listed in Table 1. The high frequency (around 100 kHz) intercept on the real axis represents the series resistance (Rs). The semicircle in the high-frequency range (100–10 kHz) results from charge-transfer resistance (Rct) at the electrolyte–counter electrode interface.33,34 The Rct value of mesoporous Ni0.85Se spheres (2.4 Ω) was smaller than that of Ni0.85Se nanoparticles (3.46 Ω). In addition, the Rct value is in a contradictory relationship with the charge-transfer ability.35,36 The lower Rct value of mesoporous Ni0.85Se spheres revealed its better charge-transfer ability in comparison with Ni0.85Se nanoparticles.
 |
| Fig. 4 Nyquist plots for symmetric cells fabricated with mesoporous Ni0.85Se spheres, Ni0.85Se nanoparticles and Pt. | |
Table 1 Corresponding EIS, Tafel polarization and CV parameters of the DSSCs assembled with mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles CEs
CEs |
Rs (Ω) |
Rct (Ω) |
Epp (mV) |
JA (mA cm−2) |
J0 log (mA cm−2) |
Jlim log (mA cm−2) |
Mesoporous Ni0.85Se spheres |
20.78 |
2.40 |
351 |
1.91 |
1.96 |
0.59 |
Ni0.85Se nanoparticles |
20.61 |
3.46 |
404 |
0.80 |
1.84 |
0.55 |
Pt |
21.60 |
1.28 |
429 |
1.81 |
2.18 |
0.85 |
In order to further evaluate the electrocatalytic performance, the cyclic voltammetry (CV) was measured in a three-electrode system. As shown in Fig. 5, all CEs have typical curves with two pairs of redox peaks and two cathodic peaks, which are related to the two-step process of I3− reduction denoted in eqn (1) and (2):
 |
| Fig. 5 Cyclic voltammograms for iodide/triiodide redox species of mesoporous Ni0.85Se spheres, Ni0.85Se nanoparticles and Pt. | |
Peak-to-peak separation (Epp) in lower potential is negatively correlated with the standard electrochemical rate constant and has positive correlation with overpotential losses.37,38 As the Fig. 5 shows, the peak-to-peak separation (Epp) values of mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles were about 351 mV and 404 mV. The Epp trend revealed the better catalytic activity for reducing I3− of mesoporous Ni0.85Se spheres than Ni0.85Se nanoparticles. As mentioned above, mesoporous structures promote the interface contact between electrode and electrolyte, and contribute to the diffusion of liquid electrolyte into the inside of mesoporous Ni0.85Se spheres. Therefore, the rate of electrocatalytic reaction the catalytic activity at the electrolyte–mesoporous Ni0.85Se spheres electrode interface was improved and faster than Ni0.85Se nanoparticles. Consequently, the peak current density (JA) of mesoporous Ni0.85Se spheres CE (1.74 mA cm−2) in lower potential is higher than that of Ni0.85Se nanoparticles CE (0.78 mA cm−2), and the JA is positively correlated with the electrocatalytic reaction rate.39,40
Fig. 6 shows the Tafel polarization curves of different CEs which were measured by the dummy cells. The polarization zone of Tafel polarization curve corresponds to the high frequency region of Nyquist plots, and diffusion zone relates to the low frequency region in EIS.26 In the polarization zone of Tafel polarization curve, as the reaction velocity of I3− reducing to I− unmatches with the transfer speed of photo-generated electrons, electrons or ions accumulate in CEs and generate the overpotential, then electrochemical polarization appears and the voltage of DSSC devices will reduce. Herein, the exchange current density (J0) in the polarization zone corresponds with Rct of Nyquist plots. Accordingly, as to diffusion zone, the inconformity between diffusion rate of iodide ions and reaction velocity of I3− to I− induces the concentration polarization. Similarly, the limiting diffusion current density (Jlim) can represent diffusion coefficient of I3− and I−. As shown in Table 1, similar with the Rct values, the calculated J0 (1.96
log (mA cm−2)) and Jlim (0.59
log (mA cm−2)) values (Fig. S5†) of mesoporous Ni0.85Se spheres were both higher than those of Ni0.85Se nanoparticles (J0 = 1.84
log (mA cm−2), Jlim = 0.55
log (mA cm−2)). J0 and Jlim both agreed well with the results of EIS experiments, and demonstrated the faster charge transfer speed and diffusion rate of mesoporous Ni0.85Se spheres than those of Ni0.85Se nanoparticles.41
 |
| Fig. 6 Tafel polarization curves for symmetric cells fabricated with mesoporous Ni0.85Se spheres, Ni0.85Se nanoparticles and Pt. | |
Current density–voltage (J–V) measurements under the AM 1.5 G condition were carried out to compare the comprehensive electrocatalytic properties of mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles. Fig. 7 shows the J–V curves for DSSCs using the above three CEs, and corresponding photovoltaic parameters of open-circuit voltage (Voc), short-circuit current (Jsc), fill factor (FF) and PCE are summarized in Table 2. Mesoporous Ni0.85Se spheres owned higher Jsc (14.40 mA cm−2), FF (0.67) and PCE (7.24%) than those of Ni0.85Se nanoparticles (Jsc = 14.00 mA cm−2, FF = 0.65, PCE = 6.73%), which demonstrated comprehensively better electrocatalytic properties of mesoporous Ni0.85Se spheres than Ni0.85Se nanoparticles. This advantage can be visually attributed to the higher current intensity of mesoporous Ni0.85Se spheres than Ni0.85Se nanoparticles, and in Fig. 7 the J–V curve of mesoporous Ni0.85Se spheres overall located above that of Pt. Simultaneously, higher FF value of mesoporous Ni0.85Se spheres, revealing their better utilization efficiency of photogenerated electron than Ni0.85Se nanoparticles.42 As mentioned above, mesoporous structures of mesoporous Ni0.85Se spheres promote the interface contact between electrode and electrolyte, and contribute to the collision among I3− and catalytic active sites. EIS, CV and Tafel polarization already demonstrated that mesoporous Ni0.85Se spheres exhibited better catalytic activity and charge transfer ability, which accelerated the rate of electrocatalytic reaction and further improved the PCE value.
 |
| Fig. 7 Photocurrent density–voltage characteristics of DSSCs with mesoporous Ni0.85Se spheres, Ni0.85Se nanoparticles and Pt. | |
Table 2 Corresponding photovoltaic parameters of the DSSCs assembled with mesoporous Ni0.85Se spheres and Ni0.85Se nanoparticles CEs
CEs |
Voc (V) |
Jsc (mA cm−2) |
FF |
PCE (%) |
Mesoporous Ni0.85Se spheres |
0.75 |
14.40 |
0.67 |
7.24 |
Ni0.85Se nanoparticles |
0.74 |
14.00 |
0.65 |
6.73 |
Pt |
0.76 |
14.80 |
0.67 |
7.56 |
Enhanced electrocatalytic performance of mesoporous Ni0.85Se spheres with RGO and SWCNT
In order to improve the electrocatalytic performance of mesoporous Ni0.85Se spheres and avert the adverse effect of the poor interdomain electron transport among spheres, introducing carbon materials (RGO or SWCNT) to take advantage of their high specific surface area and extremely high electrical conductivity should be an ideal idea.43 Therefore, RGO and SWCNT with different amount were introduced by facile physical mix to enhance electrocatalytic activity of mesoporous Ni0.85Se spheres CE. The information about RGO and SWCNT is shown in ESI.† On one hand, EIS experiments (Fig. 8a and b and S6†) were performed to measure the effect of mixing carbon materials with mesoporous Ni0.85Se spheres on charge transfer ability, and detailed EIS parameters are summarized in Table 3. As mentioned, Rct value was 2.40 Ω for Ni0.85Se, when mesoporous Ni0.85Se spheres mixing with RGO (NRCs), the Rct values of Ni0.85Se + 3 wt% RGO, Ni0.85Se + 6 wt% RGO and Ni0.85Se + 9% RGO were 0.82 Ω, 0.68 Ω and 0.86 Ω, respectively. As to Ni0.85Se + SWCNT CEs (NSCs), the mixtures of Ni0.85Se with 3 wt% (Rct = 0.93 Ω), 6 wt% (Rct = 0.90 Ω) and 9 wt% (Rct = 0.74 Ω) SWCNT also exhibited lower Rct than that of mesoporous Ni0.85Se spheres. Furthermore, all of the Rct values of NRCs and NSCs are lower than that of Pt CE (1.28 Ω). Obviously, the addition of RGO and SWCNT greatly enhanced the charge transfer ability of mesoporous Ni0.85Se spheres. In consideration of the reason, it may due to that the addition of RGO and SWCNT not only improved the charge transfer between Ni0.85Se spheres, but also promoted the transfer of photo-generated charge from CEs materials to electrolyte molecules.44,45 On the other hand, Tafel polarization and CV experiments of NRCs and NSCs were also carried out to demonstrate the enhancement effect of RGO and SWCNT on the electrocatalytic property of mesoporous Ni0.85Se spheres CE, and relevant information was listed in ESI Fig. S7–S9.†
 |
| Fig. 8 Nyquist plots (a and b) and photocurrent density–voltage characteristics (c and d) of DSSCs fabricated with NRCs, NSCs and Pt. | |
Table 3 Corresponding photovoltaic and electrochemical impedance parameters of the DSSCs assembled with various CEs
CEs |
Voc (V) |
Jsc (mA cm−2) |
FF |
PCE (%) |
Rs (Ω) |
Rct (Ω) |
Ni0.85Se + 3 wt% RGO |
0.74 |
15.60 |
0.66 |
7.58 |
20.11 |
0.82 |
Ni0.85Se + 6 wt% RGO |
0.76 |
15.20 |
0.69 |
7.87 |
20.25 |
0.68 |
Ni0.85Se + 9 wt% RGO |
0.74 |
15.20 |
0.68 |
7.62 |
19.58 |
0.86 |
RGO |
0.75 |
11.20 |
0.67 |
5.61 |
19.21 |
7.16 |
Ni0.85Se + 3 wt% SWCNT |
0.76 |
14.00 |
0.71 |
7.45 |
20.62 |
0.93 |
Ni0.85Se + 6 wt% SWCNT |
0.75 |
14.80 |
0.68 |
7.50 |
20.52 |
0.90 |
Ni0.85Se + 9 wt% SWCNT |
0.76 |
14.80 |
0.69 |
7.74 |
20.43 |
0.74 |
SWCNT |
0.77 |
11.60 |
0.68 |
6.07 |
18.32 |
6.46 |
Photoelectric conversion efficiencies of NRCs and NSCs (Fig. 8c and d and S10†) were measured to evaluate the overall electrocatalytic effect between mesoporous Ni0.85Se spheres and carbon materials. As to the mesoporous Ni0.85Se spheres CE, PCE value was 6.73%. With the addition of RGO, the photovoltaic performances (Table 3) have been improved substantially, and PCE values of Ni0.85Se + 3 wt% RGO (7.58%), Ni0.85Se + 6 wt% RGO (7.87%) and Ni0.85Se + wt% RGO (7.62%) were higher than those of mesoporous Ni0.85Se spheres and Pt (7.56%) CEs. SWCNT also evidently raise the PCE of mesoporous Ni0.85Se spheres CE, and PCE values of Ni0.85Se + 3 wt% SWCNT, Ni0.85Se + 6 wt% SWCNT and Ni0.85Se + wt% SWCNT were 7.45%, 7.50% and 7.74%, respectively. As the photovoltaic parameters in Table 3 show, NRCs and NSCs both showed higher values of FF and much higher Jsc values than those of mesoporous Ni0.85Se spheres, which represented that both of RGO and SWCNT not only promoted the charge transfer in composites but also enhanced the utilization efficiency of photo-generated charge.30
As mentioned in the above results, different amount of RGO or SWCNT exhibited different effect on the electrocatalytic property of mesoporous Ni0.85Se spheres CE. In consideration of reaction mechanism, the lower percentage of RGO or SWCNT than the ideal amount in composites cannot reach the requirement of charge-transfer, and the catalytic activity of the composites cannot get fully exploited. This condition can be demonstrated by the results of EIS and CV experiments. When the amount of RGO or SWCNT is more than the ideal amount, the overmuch RGO or SWCNT will decrease the amount of active sites in the basal plane for I3− electrocatalysis. And weak adhesion of overmuch RGO or SWCNT to conductive substrate hinders the reception of photo-generated electrons from photoanode, which is adverse to charge-transfer and catalytic reaction.46 Herein, Ni0.85Se + 6 wt% RGO and Ni0.85Se + 9 wt% SWCNT CEs obtained the best charge-transfer ability and electrocatalytic performance in NRCs and NSCs, respectively.
Conclusions
In conclusion, we have demonstrated a facile self-assembly method to fabricate mesoporous Ni0.85Se spheres consisted of numerous primary particles. Their mesoporous structures and large internal reaction area endow mesoporous Ni0.85Se spheres with excellent performance as CEs of DSSCs. It is evidenced that mesoporous Ni0.85Se spheres owned overall better electrocatalytic activity for the reduction of I3− than Ni0.85Se nanoparticles. On the other hand, the addition of RGO and SWCNT substantially improved the interdomain electron transport among Ni0.85Se spheres and accelerated the velocity of electrocatalytic reaction. Simultaneously, the most appropriate amount of carbon materials resulted in the optimal effect for enhancing the electrocatalytic performance in DSSCs.
Acknowledgements
This work was supported by the National Science Foundation of China (No. 21271108), the Ministry of Science and Technology (Grant 2014CB932001), the China-U.S. Center for Environmental Remediation and Sustainable Development, the Key Project of Chinese National Programs for Fundamental Research and Development (973 Program) (Grant 2014CB932001), National Natural Science Foundation of China (No. 21425729).
Notes and references
- A. Walcarius, Chem. Soc. Rev., 2013, 42, 4098–4140 RSC.
- M. E. Davis, Nature, 2002, 417, 813–821 CrossRef CAS PubMed.
- Y. Shi, C. Hua, B. Li, X. Fang, C. Yao, Y. Zhang, Y.-S. Hu, Z. Wang, L. Chen, D. Zhao and G. D. Stucky, Adv. Funct. Mater., 2013, 23, 1832–1838 CrossRef CAS.
- Y. Wan and D. Zhao, Chem. Rev., 2007, 107, 2821–2860 CrossRef CAS PubMed.
- J. Zeng, C. Francia, J. Amici, S. Bodoardo and N. Penazzi, J. Power Sources, 2014, 272, 1003–1009 CrossRef CAS.
- W. Li, Q. Yue, Y. Deng and D. Zhao, Adv. Mater., 2013, 25, 5129–5152 CrossRef CAS PubMed.
- M. Chen, L. L. Shao, X. Qian, T. Z. Ren and Z. Y. Yuan, J. Mater. Chem. C, 2014, 2, 10312–10321 RSC.
- Y. Zhang, C. Jiang, S. Zhuang, M. Lu and Y. Cai, RSC Adv., 2016, 6, 49782–49786 RSC.
- P. G. Bruce, B. Scrosati and J. M. Tarascon, Angew. Chem., Int. Ed., 2008, 47, 2930–2946 CrossRef CAS PubMed.
- J. S. Xu and Y. J. Zhu, ACS Appl. Mater. Interfaces, 2012, 4, 4752–4757 Search PubMed.
- F. Huang, D. Chen, Y. Chen, R. A. Caruso and Y. B. Cheng, J. Mater. Chem. C, 2014, 2, 1284–1289 RSC.
- P. N. T. Santanu Bag, P. J. Chupas, G. S. Armatas and M. G. Kanatzidis, Science, 2007, 317, 490–493 CrossRef PubMed.
- J. Xu, H. Xue, X. Yang, H. Wei, W. Li, Z. Li, W. Zhang and C. S. Lee, Small, 2014, 10, 4754–4759 CrossRef CAS PubMed.
- M. G. Kanatzidis, Adv. Mater., 2007, 19, 1165–1181 CrossRef CAS.
- D. Lin, Y. Li, P. Zhang, W. Zhang, J. Ding, J. Li, G. Wei and Z. Su, RSC Adv., 2016, 6, 52739–52745 RSC.
- W. W. Zhao, C. Zhang, F. Y. Geng, S. F. Zhuo and B. Zhang, ACS Nano, 2014, 8, 10909–10919 CrossRef CAS PubMed.
- S. Peng, T. Zhang, L. Li, C. Shen, F. Cheng, M. Srinivasan, Q. Yan, S. Ramakrishna and J. Chen, Nano Energy, 2015, 16, 163–172 CrossRef CAS.
- M. R. Gao, Y. F. Xu, J. Jiang and S. H. Yu, Chem. Soc. Rev., 2013, 42, 2986–3017 RSC.
- W. Ma, Y. Guo, X. Liu, D. Zhang, T. Liu, R. Ma, K. Zhou and G. Qiu, Chem.–Eur. J., 2013, 19, 15467–15471 CrossRef CAS PubMed.
- A. Y. Zhang, Q. Ma, M. K. Lu, G. J. Zhou, C. Z. Li and Z. G. Wang, J. Phys. Chem. C, 2009, 113, 15492–15496 CrossRef CAS.
- N. Moloto, M. J. Moloto, N. J. Coville and S. Sinha Ray, J. Cryst. Growth, 2009, 311, 3924–3932 CrossRef CAS.
- G. Q. Zhang, W. Wang, Q. X. Yu and X. G. Li, Chem. Mater., 2009, 21, 969–974 CrossRef CAS.
- H. Wang, W. Wei and Y. H. Hu, Top. Catal., 2014, 57, 607–611 CrossRef CAS.
- A. T. Swesi, J. Masud and M. Nath, Energy Environ. Sci., 2016, 9, 1771–1782 Search PubMed.
- X. Zhang, S. Guo, M. Zhen, G. Gao and L. Liu, J. Electrochem. Soc., 2015, 162, 774–779 CrossRef.
- Z. Zhuang, Q. Peng, J. Zhuang, X. Wang and Y. Li, Chem.–Eur. J., 2005, 12, 211–217 CrossRef PubMed.
- M. Z. Xue and Z. W. Fu, Electrochem. Commun., 2006, 8, 1855–1862 CrossRef CAS.
- F. Gong, H. Wang, X. Xu, G. Zhou and Z. S. Wang, J. Am. Chem. Soc., 2012, 134, 10953–10958 CrossRef CAS PubMed.
- Y. Duan, Q. Tang, J. Liu, B. He and L. Yu, Angew. Chem., Int. Ed., 2014, 53, 14569–14574 CrossRef CAS PubMed.
- X. Zhang, Y. Yang, S. Guo, F. Hu and L. Liu, ACS Appl. Mater. Interfaces, 2015, 7, 8457–8464 Search PubMed.
- H. Wang and Y. H. Hu, Energy Environ. Sci., 2012, 5, 8182–8188 Search PubMed.
- J. Di, J. Xia, Y. Ge, L. Xu, H. Xu, M. He, Q. Zhang and H. Li, J. Mater. Chem. A, 2014, 2, 15864–15874 RSC.
- Z. Q. Wan, C. Y. Jia and Y. Wang, Nanoscale, 2015, 7, 12737–12742 RSC.
- X. Ma, H. Elbohy, S. Sigdel, C. Lai, Q. Qiao and H. Fong, RSC Adv., 2016, 6, 11481–11487 RSC.
- D. Sebastián, V. Baglio, M. Girolamo, R. Moliner, M. J. Lázaro and A. S. Aricò, J. Power Sources, 2014, 250, 242–249 CrossRef.
- L. Zhang and Z. S. Wang, J. Mater. Chem. C, 2016, 4, 3614–3620 RSC.
- Y. Duan, Q. Tang, B. He, R. Li and L. Yu, Nanoscale, 2014, 6, 12601–12608 RSC.
- S. Hussain, S. F. Shaikh, D. Vikraman, R. S. Mane, O.-S. Joo, M. Naushad and J. Jung, RSC Adv., 2015, 5, 103567–103572 RSC.
- X. Zhang, X. Huang, C. Li and H. Jiang, Adv. Mater., 2013, 25, 4093–4096 CrossRef CAS PubMed.
- H. C. Chen, Y. H. Chen, C. C. Liu, Y. C. Chien, S. W. Chou and P. T. Chou, Chem. Mater., 2012, 24, 4766–4772 CrossRef CAS.
- H. Cai, Q. Tang, B. He and P. Li, J. Power Sources, 2014, 258, 117–121 CrossRef CAS.
- P. Cheng, Y. Wang, L. Xu, P. Sun, Z. Su, F. Jin, F. Liu, Y. Sun and G. Lu, RSC Adv., 2016, 6, 51320–51326 RSC.
- J. Xie, J. Zhang, S. Li, F. Grote, X. Zhang, H. Zhang, R. Wang, Y. Lei, B. Pan and Y. Xie, J. Am. Chem. Soc., 2013, 135, 17881–17888 CrossRef CAS PubMed.
- Z. Yin, J. Zhu, Q. He, X. Cao, C. Tan, H. Chen, Q. Yan and H. Zhang, Adv. Energy Mater., 2014, 4, 1300574–1300592 CrossRef.
- M. Y. Liu, G. Li and X. S. Chen, ACS Appl. Mater. Interfaces, 2014, 6, 2604–2610 Search PubMed.
- X. H. Miao, K. Pan, G. F. Wang, Y. P. Liao, L. Wang, W. Zhou, B. J. Jiang, Q. J. Pan and G. H. Tian, Chem.–Eur. J., 2014, 20, 474–482 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra11085a |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.